arxiv: v1 [astro-ph.ep] 30 Jan 2013

Similar documents
Dynamically Unstable Planetary Systems Emerging Out of Gas Disks

Importance of the study of extrasolar planets. Exoplanets Introduction. Importance of the study of extrasolar planets

2 Ford, Rasio, & Yu. 2. Two Planets, Unequal Masses

Kozai-Lidov oscillations

Architecture and demographics of planetary systems

Kepler Planets back to the origin

Lecture Outlines. Chapter 15. Astronomy Today 7th Edition Chaisson/McMillan Pearson Education, Inc.

Testing Theories of Planet Formation & Dynamical Evolution of Planetary Systems using Orbital Properties of Exoplanets

Planet formation in protoplanetary disks. Dmitry Semenov Max Planck Institute for Astronomy Heidelberg, Germany

Data from: The Extrasolar Planet Encyclopaedia.

arxiv:astro-ph/ v2 15 May 2008

arxiv: v2 [astro-ph.ep] 30 Nov 2013

Planetary System Stability and Evolution. N. Jeremy Kasdin Princeton University

The Long-Term Dynamical Evolution of Planetary Systems

arxiv: v1 [astro-ph.ep] 3 Apr 2018

The Use of Transit Timing to Detect Extrasolar Planets with Masses as Small as Earth

A REGION VOID OF IRREGULAR SATELLITES AROUND JUPITER

Ruth Murray-Clay University of California, Santa Barbara

Lecture Outlines. Chapter 15. Astronomy Today 8th Edition Chaisson/McMillan Pearson Education, Inc.

Solar System evolution and the diversity of planetary systems

arxiv: v1 [astro-ph.ep] 20 Apr 2014

Global models of planetary system formation. Richard Nelson Queen Mary, University of London

Formation of Planets around M & L dwarfs

Internal structure and atmospheres of planets

arxiv: v4 [astro-ph.ep] 1 Aug 2013

III The properties of extrasolar planets

Lecture 20: Planet formation II. Clues from Exoplanets

Observational constraints from the Solar System and from Extrasolar Planets

Definitions. Stars: M>0.07M s Burn H. Brown dwarfs: M<0.07M s No Burning. Planets No Burning. Dwarf planets. cosmic composition (H+He)

II. Results from Transiting Planets. 1. Global Properties 2. The Rossiter-McClaughlin Effect

Exoplanets: a dynamic field

Characterization of the exoplanet host stars. Exoplanets Properties of the host stars. Characterization of the exoplanet host stars

Dynamic Exoplanets. Alexander James Mustill

Habitability in the Upsilon Andromedae System

EART164: PLANETARY ATMOSPHERES

Observations of extrasolar planets

Chapter 15 The Formation of Planetary Systems

Other Planetary Systems (Chapter 13) Extrasolar Planets. Is our solar system the only collection of planets in the universe?

The dynamical evolution of exoplanet systems. Melvyn B. Davies Department of Astronomy and Theoretical Physics Lund University

arxiv:astro-ph/ v1 25 Aug 1998

The stability of planets in the Alpha Centauri system

Observational Cosmology Journal Club

What can be learned from the dynamics of packed planetary systems?

Accretion of Planets. Bill Hartmann. Star & Planet Formation Minicourse, U of T Astronomy Dept. Lecture 5 - Ed Thommes

MINIMUM MASS EXTRASOLAR NEBULA DERIVED FROM THE MASS OF EXTRASOLAR PLANETARY SYSTEMS

Astronomy 405 Solar System and ISM

The dynamical evolution of transiting planetary systems including a realistic collision prescription

What Have We Found? 1978 planets in 1488 systems as of 11/15/15 ( ) 1642 planets candidates (

Possible commensurabilities among pairs of extrasolar planets

Class 15 Formation of the Solar System

PLANETARY FORMATION THEORY EXPLORING EXOPLANETS

Life in the Universe (1)

EXOPLANET LECTURE PLANET FORMATION. Dr. Judit Szulagyi - ETH Fellow

Hydrodynamic Outcomes. Transitional Discs. of Planet Scattering in. Nick Moeckel IoA. Phil Armitage Colorado

Migration. Phil Armitage (University of Colorado) 4Migration regimes 4Time scale for massive planet formation 4Observational signatures

Astronomy. physics.wm.edu/~hancock/171/ A. Dayle Hancock. Small 239. Office hours: MTWR 10-11am

Dynamical properties of the Solar System. Second Kepler s Law. Dynamics of planetary orbits. ν: true anomaly

arxiv: v2 [astro-ph.ep] 10 Feb 2015

Exoplanet Mass, Radius, and the Search for Habitable Worlds

On the Origin of the Rocky Planets, Fugue in Venus Megacollision

FORMING DIFFERENT PLANETARY ARCHITECTURES. I. FORMATION EFFICIENCY OF HOT JUPITES FROM HIGH-ECCENTRICITY MECHANISMS

Extrasolar Planets. Properties Pearson Education Inc., publishing as Pearson Addison-Wesley

THE PLANE OF THE KUIPER BELT

Gas Disks to Gas Giants: Simulating the Birth of Planetary Systems

The exoplanet eccentricity distribution from Kepler planet candidates

Formation of the Solar System Chapter 8

How do we model each process of planet formation? How do results depend on the model parameters?

What is it like? When did it form? How did it form. The Solar System. Fall, 2005 Astronomy 110 1

Planetary system dynamics Part III Mathematics / Part III Astrophysics

arxiv: v1 [astro-ph.ep] 15 Apr 2016

Formation of the Solar System. What We Know. What We Know

Why Search for Extrasolar Planets?

Joseph Castro Mentor: Nader Haghighipour

Minimum Radii of Super-Earths: Constraints from Giant Impacts

Dynamical Stability of Terrestrial and Giant Planets in the HD Planetary System

Young Solar-like Systems

Practical Numerical Training UKNum

Circumbinary Planets/Disks

Other planetary systems

THE CURIOUSLY WARPED MEAN PLANE OF THE KUIPER BELT

Eccentricity pumping of a planet on an inclined orbit by a disc

Chapter 15: The Origin of the Solar System

Probing the Galactic Planetary Census

arxiv: v1 [astro-ph] 23 May 2007

arxiv: v1 [astro-ph.ep] 27 Sep 2016

Origin of the Solar System

Gravitational microlensing. Exoplanets Microlensing and Transit methods

Secular Planetary Dynamics: Kozai, Spin Dynamics and Chaos

Who was here? How can you tell? This is called indirect evidence!

AST111, Lecture 1b. Measurements of bodies in the solar system (overview continued) Orbital elements

The Physics of Exoplanets

The Transit Method: Results from the Ground

FORMATION OF HOT PLANETS BY A COMBINATION OF PLANET SCATTERING, TIDAL CIRCULARIZATION, AND THE KOZAI MECHANISM

Tidal Dissipation in Binaries

18 An Eclipsing Extrasolar Planet

RING DISCOVERED AROUND DWARF PLANET

Planetary Systems in Stellar Clusters

9. Formation of the Solar System

Discovering Exoplanets Transiting Bright and Unusual Stars with K2

Astronomy 241: Review Questions #2 Distributed: November 7, 2013

Transcription:

Draft version June 27 2018 Preprint typeset using L A TEX style emulateapj v. 5/2/11 TESTING IN SITU ASSEMBLY WITH THE KEPLER PLANET CANDIDATE SAMPLE Brad M. S. Hansen 1 & Norm Murray 23 Draft version June 27 2018 arxiv:1301.7431v1 [astro-ph.ep] 30 Jan 2013 ABSTRACT We present a Monte Carlo model for the structure of low mass (total mass < 25M ) planetary systems that form by the in situ gravitational assembly of planetary embryos into final planets. Our model includes distributions of mass eccentricity inclination and period spacing that are based on the simulation of a disk of 20M forming planets around a solar mass star and assuming a power law surface density distribution a 1.5. The output of the Monte Carlo model is then subjected to the selection effects that mimic the observations of a transiting planet search such as that performed by the Kepler satellite. The resulting comparison of the output to the properties of the observed sample yields an encouraging agreement in terms of the relative frequencies of multiple planet systems and the distribution of the mutual inclinations when moderate tidal circularisation is taken into account. The broad features of the period distribution and radius distribution can also be matched within this framework although the model underpredicts the distribution of small period ratios. This likely indicates that some dissipation is still required in the formation process. The most striking deviation between model and observations is in the ratio of single to multiple systems in that there are roughly 50% more single planet candidates observed than are produced in any model population. This suggests that some systems must suffer additional attrition to reduce the number of planets or increase the range of inclinations. Subject headings: planet-star interactions; planets and satellites: dynamical evolution and stability 1. INTRODUCTION The search for planets around stars other than the Sun has revealed a great diversity of both planet mass and location. Jovian mass planets have been discovered both in very short period orbits (Mayor & Queloz 1995 Butler et al. 1997) at distances more characteristic of the planets in our own solar system (Butler & Marcy 1996; Boisse et al. 2012) at great distances from their host stars (Udalski et al. 2005; Gaudi et al. 2008; Kalas et al. 2008; Marois et al. 2008 2010) and potentially even as free floating objects (Sumi et al. 2011). Lower mass planets are harder to detect but planets in the Uranus/Neptune mass range (10M 100M ) have also been found with a range of semimajor axes both close in (Howard et al. 2010; Mayor et al. 2011; Borucki et al. 2010; Holman et al. 2010; Lissauer et al. 2011a) and at large separations (Gould et al. 2006 2010; Sumi et al. 2010). Furthermore multiple detection techniques are now able to probe below 10M (Wolszczan & Frail 1992; McArthur et al. 2004; Santos et al. 2004; Rivera et al. 2005; Beaulieu et al. 2006; Lovis et al. 2006; Charbonneau et al. 2009; Queloz et al. 2009; Mayor et al. 2009; Batalha et al. 2011) potentially delving into the regime of true terrestrial planet formation. Indeed the number of planet candidates emerging from the Kepler satellite data suggests that this class may soon dominate the demographics of known extrasolar planets (Borucki et al. 2011). The origin of these different classes of planet is still a matter of active discussion. The presence of so-called hot Jupiters close to their parent stars was initially held to be securely described within the context of planetary migration (Lin Bodenheimer & Richardson 1996) in which giant planets form at distances several AU and migrate inwards due to torques from the gaseous protoplanetary disk (Goldreich & Tremaine 1980; Ward 1997). The discovery of significant misalignments between the planetary orbital angular momenta and the stellar star host spin (Winn et al. 2010 and historical references therein) has cast some doubt on the applicability of this model to at least some of the observed systems reviving interest in such mechanisms as planetary scattering (Rasio & Ford 1996; Nagasawa & Ida 2011) or Kozai migration (Wu & Murray 2003; Fabrycky & Tremaine 2007; Naoz et al. 2011). Another significant problem for the migration paradigm is the fact that both radial velocity and transit methods find Neptune mass planets in a part of parameter space that such models predicted to be devoid of planets (Ida & Lin 2008). This does not necessarily imply that migration does not occur but it does indicate that current models are either not universally applicable or are missing one or more crucial elements. Such advances have spurred new interest in the origin of low mass planets around other stars and generated alternative proposals. It is possible that such planets did migrate inwards but were subject to additional perturbations either from mutual gravitational interactions (Terquem & Papaloizou 2007; Ida & Lin 2010) or some other stochastic forcing (Rein 2012) in order to match the observed period ratio distribution in multiple systems. Alternatively it has 1 Department of Physics & Astronomy and Institute of Geophysics & Planetary Physics University of California Los Angeles Los Angeles CA 90095 hansen@astro.ucla.edu 2 Canadian Institute for Theoretical Astrophysics 60 St. George Street University of Toronto Toronto ON M5S 3H8 Canada;murray@cita.utoronto.ca 3 Canada Research Chair in Astrophysics

2 Kepler Rocks been proposed that intense stellar irradiation can evaporate the envelopes of hot Jupiter planets leaving behind only the Neptune-mass stellar core (Baraffe et al. 2004). However current estimates of the evaporation rates for known Jupiter planets (Ehrenreich & Desert 2011) suggest that these do not compose a viable progenitor population. Instead Hansen & Murray (2011) propose that this population is the result of in situ assembly of rocky planets which may capture substantial but not dominant gaseous envelopes if they grow to sufficient size before the gaseous disk disperses. This proposal has the advantage that it naturally produces multiple planet systems as are observed and furthermore matches the statistical properties of the observed multiple planet systems. The original intent of the Hansen & Murray (2011) proposal was to match the planets observed in the radial velocity surveys descibed in Howard et al. (2010) and Mayor et al. (2011) which represent the tip of the iceberg of the low mass planet sample by virtue of the detection limits inherent in the RV method. However the Kepler mission (Borucki et al. 2011) has begun to expose more of the iceberg suggesting that there are many systems with lower mass planets offering the possibility of further testing the in situ assembly model. In particular these lower mass planets most likely did not substantially increase their masses by accreting gas from the nebula and so should represent a more straightforward test of the model without uncertainties in the treatment of gas accretion. Therefore in this paper we perform a quantitative comparison of in situ assembly models with a planetary candidate sample culled from the Kepler satellite data under the assumption that these are primarily rocky planets and that their organization reflects the conditions that emerge directly from the gravitational assembly of the planetary system. In 2 we describe our model for the original rocky protoplanetary disk and its assembly into the final planetary configuration. We also develop statistical descriptions for the distribution of various important quantities such as planetary masses and mutual inclinations which are then used to simulate the distribution of a large population of planetary systems in a Monte Carlo fashion and to quantify their observability when studied in the manner of a search for transits. These are then compared to a well-defined subset of the Kepler candidates in 3 and the implications are discussed in 4. 2. PLANETARY ASSEMBLY SIMULATIONS Hansen & Murray (2011) hereafter HM11 simulated the assembly of planets for a variety of rocky disks with masses ranging up 100M within 1 AU and accounting for both purely rocky systems and for the accretion of substantial amounts of gas from the surrounding nebula. This was dictated by a desire to match systems detected in radial velocity surveys whose masses suggest rock-to-gas ratios of order unity (and whose existence is born out by a handful of transiting systems such as HD149026b HAT-P-11 and Kepler-9bc). However statistical analyses of the Kepler sample (e.g. Howard et al. 2012 Youdin 2011) suggest that this population contains a substantial contribution from bodies whose mass is predominantly rocky even if their radii are inflated by a minority component (by mass) of gas. Therefore this new sample by virtue of its size and different detection technique offers a new test of planetary origin models. Our model for the origin of these planetary systems is that the final assembly stage from planetary embryos to final planets occurred in situ with the masses and separations of the final planets determined by the gravitational interactions between the assembling bodies. Whether the planetary embryos are formed from rocky material that condensed out of the gas in situ or migrated inwards as small bodies is not specified and may be a function of total disk mass. In HM11 the disk masses assumed were in excess of 30M and most likely did require an enhancement over the heavy element inventory of the original local gas disk. However for lower mass disks such inward migration may not be necessary. Our model for the origin of the observed Kepler systems has a surface density profile Σ a 1.5 normalised to contain 20M of rocky material interior to 1 AU around a 1M star. This is chosen to be a lower mass version of the density profile that best fit the observations in HM11 and the overall mass normalisation was chosen to be the maximum that would not result in many of the assembling protoplanets crossing the gas capture threshold defined in HM11 within 1 Myr. Interestingly these criteria yield almost exactly the same disk as has been recently derived by Chiang & Laughlin (2012) using an analogue of the Minimum Mass Solar Nebula argument but applied to the observed Kepler candidate sample. This agreement suggests an encouraging self-consistency to our adoption of this as the underlying basis for forming a Kepler sample dominated by rocky planets. Our initial conditions are chosen using the formalism of Kokubo & Ida (1998) to assign initial masses and semi-major axes for the protoplanetary embryos at the end of the oligarchic accumulation stage and are integrated forwards using the Mercury integrator (Chambers 1999) on the UCLA Hoffman2 Shared Computing Cluster. We perform 100 realisations of this model. The inner edge of the original disk is taken to be at 0.05 AU and so the time step for the integrations is only 12 hours in order to resolve the innermost orbits. The integrations are run for 10 Myr. This proves sufficient to achieve dynamical stability as evidenced both by inspection of the individual evolutionary histories and the integration of a subsample of 20 to ages of 100 Myr without further significant change in the overall configuration. This assembly time scale is shorter than has been found for the solar system terrestrial planets (e.g. Chambers & Wetherill 1998; Chambers 2001) because of the higher mass and more compact nature of the disk. We then examine the ensemble of surviving systems to characterise various properties of the resulting population. The parameters of the final systems are shown in Table 1 including a variety of statistical measures collected together by Chambers (2001) for the purposes of characterising simulations of late stage planet assembly in the Solar System. Not surprisingly the simulations here are both more closely packed (because of the larger mass in the disks relative to the solar system case studied by Chambers) and have the mass distributed over a greater range of radii (because we consider disks extending down to 0.05 AU). Furthermore many of these measures are based on the distribution

Hansen & Murray 3 Fig. 1. The solid histogram shows the distribution of planetary multiplicity for one hundred realisations of our scenario after 10 Myr (counting planets for which a < 1.1AU). The dotted histogram is the best fit Poisson distribution for the number of planets in excess of two but it is not sufficiently narrowly peaked (χ 2 = 3.9 per degree of freedom for bins N p=3 8). The dashed histogram is the squared Poisson distribution discussed in the text which is a much better fit (χ 2 = 1.1 per degree of freedom). Fig. 2. The solid histogram is the distribution of final planet masses normalised by the local value of σ M for planets with a < 1.1 AU. The dotted line is a Rayleigh distribution in the scaled variable M/σ M. of mass a quantity only indirectly accessible for the bulk of the Kepler sample. As such we will use our simulation results to build a Monte Carlo model to perform a more direct comparison with the observables of a Kepler-style experiment. 2.1. Multiplicity One of the most fundamental questions one can ask about the Kepler data set is how many planets a given system is likely to hold. Kepler has detected 361 stars (Fabrycky et al. 2012) which exhibit multiple transiting planets with the record for the largest number of transiting planets around a single star currently held by Kepler-11 (Lissauer et al. 2011a) at six. Of course the mutual inclinations of a set of planets orbiting a star will dictate what fraction transit and therefore the transiting planets represent only a subsample of the true underlying system. Figure 1 shows the distribution of true multiplicity for our simulated ensemble in which we count all planets that survive with semi-major axes < 1.1 AU (this excludes a handful of surviving embryos on eccentric orbits scattered out of the original disk) irrespective of mass. The median number of surviving planets so defined is four. Statistical analyses of Kepler data with pre-defined functions often assume a functional form for the multiplicity that is based on the Poisson distribution (e.g. Fang & Margot 2012). We find that this is not a good fit to our simulations because it is too broad (i.e. it contains too many systems with either fewer or more planets than the median value). Furthermore the minimum number of planets found in our simulations is three so that our function should be adopted to describe the number in excess of two. To that end we find that a better empirical model of the true distribution is to square the Poisson function for the number of planets in excess of two i.e. p(n) = λ2(n 2) e 2λ ((N 2)!) 2 (1) where λ = 3.8 yields the best fit. The reason for the minimum multiplicity is related to the requirements of conserving mass energy and angular momentum while assembling an extended distribution of mass into a handful of planets through mutual gravitational perturbations. We examine this issue further in 2.3.1. 2.2. Mass and Radius Distributions

4 Kepler Rocks Fig. 3. The solid points are masses and radii for extrasolar planets measured through transits and radial velocities based on the known planet sample as of Dec 2012. The open points represent the solar system planets Venus Earth Uranus and Neptune. The solid and dotted curves are theoretical mass-radius relations from Seager et al. (2007) for Perovskite and Water respectively. The long dashed line is the empirical mass-radius relationship from Lissauer et al. (2011b) and the short dashed line is the relationship derived by Tremaine & Dong (2012). Our radius estimates are based on the solid curve with a potential multiplicative factor R applied to account for potential Hydrogen envelopes. Fig. 4. The solid histogram shows the distribution of for the planet pairs with a < 1.1AU. The vertical dotted line is the stability limit for Jovian mass pairs derived by Gladman (1993). The vertical dashed line is for = 9 derived by Smith & Lissauer (2009) for earth-mass systems. We see that our distribution cuts off completely at this value although the more accurate effective cutoff is = 13 and the bulk of the pairs are more widely separated than this peaking around = 30 with few systems with > 50. The masses of surviving planets exhibit a trend of increasing mass with semi-major axis. Thus we fit the mass distribution as a Rayleigh distribution with a dispersion that depends on semi-major axis namely and f(m) = m σ 2 me 1 2 (m/σm)2 (2) σ m = 7M (a/1au) 0.6. (3) This resulting fit is shown in Figure 2. For comparison to the Kepler sample we need to convert masses to radii. In the simplest incarnation our model assumes that the masses of the planets are dominated by the rocky component. Therefore we use the mass-radius relation of Seager et al. (2007) based on the Perovskite equation of state to characterise our planetary radii. Figure 3 shows the current sample of planets with both mass and radius measurements. It is clear that many low mass planets have radii in excess of that expected from our no-gas model suggesting that the observed Kepler planet possess atmospheres of lower molecular weight material either water or gas. Previous authors have attempted to address this issue by interpolating between rocky planets and more massive gaseous planets. Lissauer et al. (2011b) suggest a mass radius relation M = M (R/R ) 2.06 based on an interpolation between the solar system terrestrial planets and ice giants. Tremaine & Dong (2012) on the other hand interpolate across the mass range probed by transiting planets. Wu & Lithwick (2012a) find an empirical mass-radius relation M = 3M (R/R ) by placing mass constraints on planetary pairs using transit-timing variations. In each case these relations amount to an implicit but not well characterized assumption about the nature of the planetary atmosphere. We adopt a radius based on the Seager et al. relationship but allow for a multiplicative radius enhancement factor R to account for a potential gaseous atmosphere. This is a reasonable approximation for the radius of a planet with a rocky core and hydrogen atmosphere of constant mass fraction at least in the regime (gas mass fractions 15% or less) necessary for the known exoplanets with masses < 10M. A pure water planet is also well described in this way with an enhancement factor R = 1.3. 2.3. Spacing of Planets

Hansen & Murray 5 The spacing of planets in multiple systems potentially contains vital information about how the system was assembled. In HM11 we examined a variety of statistics commonly used in assessing models of solar system formation and found that the in situ assembly model was able to reproduce the properties of the systems observed by Howard et al and Mayor et al. Table 1 shows the same quantities for these simulations. One of the most common measures for planetary spacing is the orbital spacing between two planets measured in terms of their average Hill radius = 2 a 2 a 1 a 2 +a 1 ( ) 1/3 m1 +m 2 where a 1 m 1 and a 2 m 2 are the semi-major axes and masses of the inner and outer planets and M is the mass of the host star. Figure 4 shows the ensemble distribution of in our simulations. The distribution is broad peaking at 30. Various authors have evaluated the lower limit on imposed by dynamical stability with > 9 being the expectation for earth-mass planets on initially circular orbits (Smith & Lissauer 2009). Our simulations do show a handful of systems just above this limit but suggest that the more significant edge to the distribution lies at 13. At the other end we find that > 50 is rare suggesting that transiting planet pairs with larger values likely host additional as yet undetected planets. The statistical measures described in HM11 and also are based on knowledge of the planetary mass. For most Kepler candidates however we have only radius measurements. Thus a more appropriate measure for the compactness of these systems is the distribution of period ratios in neighbouring pairs. Lissauer et al. (2011b) and Fabrycky et al. (2012) have discussed this quantity and demonstrated that the Kepler sample shows a broad distribution with additional structure near commensurabilities. Figure 5 shows the distribution from our simulations peaking at ratios just about two. There is some hint of structure near the 2:1 commensurability but with neither the strength or the asymmetry seen in the observational sample. Lissaueret al. (2011b)alsodefine the ζ i statistic to measurethe proximitytocommensurabilitiesoforderi. Figure6 shows the distributions of ζ 1 and ζ 2 for our simulations. We see that there is some measure of avoidance of first order resonance in the simulations and no corresponding feature with respect to second order resonance. However we again see a symmetric structure rather than the asymmetry observed in Lissauer et al or Fabrycky et al. Nevertheless the broader features of the distribution are close enough that they may provide the initial basis on top of which additional dynamical structure is created by dissipative effects such as tidal evolution as proposed by Wu & Lithwick (2012b) or Batygin & Morbidelli (2013). We can cast the same information in the direct form of a distribution of separations which we call δ (since it is essentially the un-normalised version of ) shown in Figure 8. In this case we also want to account for a slight trend of δ with semi-major axis fitting the distribution about this. The origin of the trend is the fact that annular spacing is dependant on the amount of mass in the disk as discussed in 2.3.1. Thus we normalise δ to δ 0 where ( a ) δ 0 = 0.43 0.23log 10 (4) 1AU and model the resulting distribution of ǫ = δ/δ 0 as 3M f(ǫ) e 1 2 ((ǫ 0.82)/0.07)2 +0.6e 1 2 ((ǫ 1.28)/0.12)2 +0.48e 1 2 ((ǫ 0.53)/0.1)2 +0.56e 1 2 ((ǫ 0.98)/0.05)2. (5) This is the function we will use to determine the spacing of our planetary systems. However since the mass and separations are chosen seperately we will include an additional step of removing systems from the Monte Carlo model that violate the requirements 13 < < 50. 2.3.1. An approximate theory of planet spacings It is of interest to understand how the spacing and multiplicity of these systems is determined. The underlying process is the gravitational assembly of an extended mass distribution regulated by both the requirements of global mass energy and angular momentum conservation as well as the requirement that the assembled bodies are sufficiently massive to perturb and eventually accrete any material nearby. The global conservation requirements are simple. Although gravitational perturbations can potentially eject material from a planetary system the systems we simulate are sufficiently deep in their stellar potential wells that the amount of material lost via this mechanism is negligible and similarly for accretion onto the star. Thus for a disk with surface density Σ = Σ 0 (r/r 0 ) α spread from r in to r out the total mass energy and angular momentum budgets are ( r0 2 (rout /r 0 ) 2 α (r in /r 0 ) 2 α) M tot =2πΣ 0 2 α ( r 0 (rout /r 0 ) 1 α (r in /r 0 ) 1 α) E tot = πgm Σ 0 1 α L tot =2π(GM ) 1/2 r 5/2 ( 0 (rout /r 0 ) 5/2 α (r in /r 0 ) 5/2 α) Σ 0 5/2 α (6) (7) (8) (9)

6 Kepler Rocks Fig. 5. The observed distribution of period ratios between neighhouring pairs (in which both lie within a < 1.1AU). The dotted lines indicate the positions of the 2:1 and 3:2 orbital commensurabilities. Fig. 6. The upper panel shows period ratios now binned in terms of ζ 1 the proximity to the nearest first order commensurability. Exact resonance is at ζ 1 = 0 (shown by the vertical dashed line). The surviving systems appear to avoid orbits too close to resonance. The lower panel shows ζ 2 (proximity to a second order resonance) for those systems for which ζ 1 > 1. We see little indication of structure in this instance. If this were to be incorporated into a single final planet of mass M tot then energy conservation dictates that the final semi-major axis is ( 1 α (rout /r 0 ) 2 α (r in /r 0 ) 2 α) a = r 0 2 α ((r out /r 0 ) 1 α (r in /r 0 ) 1 α (10) ) and angular momentum conservation requires that the eccentricity be given by 1 e 2 = (2 α) 3 (1 y 1 α )(y 5/2 α 1) 2 (α 1)(5/2 α) 2 (y 2 α 1) 3 (11) where y = r out /r in. It proves impossible to satisfy these two conditions unless y = 1 (i.e. unless we specify an infinitely narrow ring i.e. a single planet) for α > 0. This is because each annulus of the disk is assumed to be on a circular orbit and therefore carries the maximum angular momentum per unit energy. The energy integral is a harmonic average which shifts the final location for the putative single planet inside the median radius of the disk resulting in an excess of angular momentum per unit energy if one assumes all the mass is accumulated into a single planet. Thus we require at least two planets to conserve all of the mass energy and angular momentum simultaneously in the assembly process. In addition this assembly requires that the final planets be massive enough to gravitationally perturb and accrete all the material in the annulus into one or other of the final planets. If the original disk is not sufficiently massive the final state may naturally consist of additional planets. We can estimate this at least for the size of disk we simulate by noting that our simulations suggest that 50 is an upper limit to the degree of separation of final planets. If an annulus produces two planets at the inner and outer edges of the annulus and their mutual > 50 then it is likely that a third intermediate planet forms from material that is too far from either of the other two planets to be perturbed onto a crossing orbit. We can express as = 73.7 y 1 y+1 ( Mtot y 1/2 1 ( 20M 3.472 r 0.05AU ) 1/2 ) 1/3 (12) where r and y are now running variables representing the inner edge and width of a particular annulus corresponding to the feeding zone of a particular final planet. The number 3.472 represents the quantity y 1/2 1 for the entire original disk from 0.05 AU to 1 AU and M tot represents the mass of the entire disk normalised by a host star of mass 1M. We have also assumed α = 3/2 for this case.

Hansen & Murray 7 Fig. 7. The solid points represent Kepler planetary candidate pairs plotted as period ratio versus the orbital period of the inner member of the pair. Coloured points and chains represent successive planets in the same system. The systems shown are all ones with one or more confirmed or validated planet Kepler-11 (red); Kepler-20 (blue); Kepler-33 (green) and Kepler-45 (cyan). The curved dotted lines indicate curves of constant calculated assuming the relation in equation (12). The horizontal dashed lines indicate the period ratios associated with the 2:1 and 3:2 commensurabilities. Fig. 8. The distribution of true (i.e. not weighted by mass) spacings of neighbouring planets shows several broad peaks. The distribution shown here has had a mean trend divided out. The dashed line shows the fitting function we use to characterise this distribution. Thus as an example assuming that one planet forms at the inner edge of the disk at 0.05 AU the requirement that < 50 suggests y 2 under these conditions so that the second planet lies no further than r = 0.1 AU. Using this location as the inner edge of the next annulus suggests y 2.4 and a third planet at 0.24 AU. Similarly this indicates y 3.5 and a fourth planet at 0.85 AU which is close to the outer edge of our disk. Thus we expect at least four planets to form from our original disk and the simulations do indeed produce mostly four and five planet systems. If we repeat this exercise with 30 (closer to the median of the distribution) the values of y range from 1.5 2 and one can fit 6 planets between 0.05 AU and 1 AU and 7 if the outer edge is 1.2 AU. Thus stability considerations suggest that most systems should have between 4 and 7 planets as is indeed observed. Note that the range of y encompassed by these arguments does allow for the presence of 2:1 period ratios. Even ratios in the 3:2 range are dynamically allowed although they require values between the median and the stability limit. One can convert the quantity y into a period ratio for each planet pair. Figure 7 shows the distribution of period ratios as a function of inner planet period for all planets in multiple systems that fall within our sample cuts in 3. Also shown are lines of constant using equation (12). This is another way of representing the distribution shown in Figure 4 and we find that indeed the bulk of the planets cluster in the range 20 < < 40 and that the limits at 10 and 50 are good delimiters of the edge of the distribution. Note that the observed data are tabulated only using periods so that the uncertainties in the mass-radius relation are not relevant. Figure 7 also shows coloured chains for a handful of confirmed high multiplicity Kepler systems. We see that the outer pair of the Kepler-20 system (blue) has a period ratio that puts it in the high wing of the distribution suggesting that an additional planet may be waiting to be found between Kepler-20f and Kepler-20d. Although the period spacing between Kepler-11f and Kepler-11g deviates somewhat from that of the inner planets in the system it is still well within the expected distribution so that we do not necessarily have a strong motivation to expect an additional planet in that gap. 2.4. Eccentricities and Inclinations Our initial conditions place all the planetary embryos on circular orbits but the process of assembly generates eccentricities by both secular interactions and direct scattering and collision (e.g. Chambers & Wetherill 1998). The resulting distribution of planetary eccentricities is given by Figure 9. We see that the tail of large eccentricities is somewhat larger than expected for a Rayleigh distribution (the normal assumption for the functional form). A better fit is to replace the Gaussian function in the fitting formula with a simple exponential so that f(e) = e σe 2 e (1 e/σe (1+ 1σe ) e 1/σe ) 1 (13)

8 Kepler Rocks Fig. 9. The solid points are eccentricities for all surviving planets interior to 1.1 AU. The dotted line is our exponential Rayleigh distribution with σ e = 0.055 giving a mean eccentricity of 0.11. Fig. 10. We show here the inclinations relative to the original orbital plane of all surviving planets in our 100 realisations of the model. We see that the range of inclinations at semi-major axis < 0.1AU is somewhat larger than those for 0.1AU < a < 1AU so we will fit to those two distributions seperately. with σ e = 0.055. This results in a mean eccentricity of 0.11 broadly consistent with the analysis of Moorhead et al. (2011) based on the distribution of Kepler transit durations. The underlying eccentricity distribution will also be affected by tidal circularisation for those systems with sufficiently small semi-major axis. In the absence of a more detailed model for the dissipation in these low mass planets we can adopt the traditional parameterisation and estimate the circularisation semi-major axis as a circ = 0.15AU ( T 5Gyr ) 2/13 ( Rp 2R ) 10/13 ( M 5M ) 2/13 ( Q 100 ) 2/13 (14) where T is the age of the host strs R p and M p are the masses of the planets and Q is the quality factor of the dissipation in the planet. Thus values of Q 10 1000 will place a circ = 0.1 0.2 AU. The inclinations of the planetary orbits are of particular interest in the case of Kepler comparisons since they will determine what fraction of a system s planets actually transit the host star from a given observational direction. The ensemble of inclinations (relative to the original orbital plane) is shown in Figure 10. We see that there is a trend for larger inclinations at small semimajor axis so we model this as two separate inclination distributions shown in Figure 11. For semi-major axes a < 0.1AU we model the distribution as f 1 (i) ie i/4 (15) while for a > 0.1AU we model it as f 2 (i) i [ ] 2 (i/0.9 ) 2 i 3.9 2 +0.35e 1 2 0.9 +0.008 i 2 (i/10.3 ) 3. (16) 1 e 1 1 e 1 We see that this latter distribution contains a component corresponding to a Rayleigh distribution with a dispersion similar to those determined by previous analyses of the Kepler data (e.g. Lissauer et al. 2011b Fabrycky et al. 2012 Tremaine & Dong 2012 Fang & Margot 2012). However it also contains a higher inclination component which may bias the inferences based on an assumed distribution of the Rayleigh form. Similarly the inclination distribution for the closest planets has a larger tail to high inclinations than that of a Rayleigh distribution. The inclination distribution describes the orientation of the orbital plane relative to that of the original disk. In order to fully describe the orientation relative to a given observer we also need to specify the angle of the line of nodes. Figure 12 shows the offsets in this angle Ω for neighbouring pairs of planets calculated in the original orbital plane of the disk. We see that the orientations of neighbouring orbits are not completely uncorrelated and that there is a tendency to avoid orbits that are exactly aligned with each other. We model this with the simple piecewise distribution shown in Figure 12.

Hansen & Murray 9 Fig. 11. The points in the upper panel show the distribution of inclinations (relative to the original orbital plane) of all planets inside 0.1 AU. The points in the lower panel show the same quantity for 0.1AU < a < 1.1AU. The dashed lines show the models described in the text. Fig. 12. The binned data show the difference in the longitude of the ascending node for neighbouring pairs of planets. There is a slight avoidance of orbits that are too close to aligned and a slight preference for Ω 90 160. We model this using the simple distribution given by the dashed line. 2.5. Monte Carlo Model Kepler has detected thousands of planets in several hundred systems which represent a sampling of a potentially much larger parameter space. Simulating such a data set directly is unrealistic but we can use the parameter distributions of the previous sections to construct a Monte Carlo model for the generation of planetary systems. This will produce model populations with the same properties as the simulations except for the exclusion of any parameter correlations (such as the mass-period trend and nodal distribution) not explicitly identified. We draw planetary separations from the distribution function f(ǫ) and planetary masses from f(m) (out to distances of 1 AU). We then evaluate the value of for each pair and reject any systems containing a pair that violates < 13 or > 50 (this is a small minority of cases because the distribution of f( ) is closely related to f(ǫ) and f(m)). We also draw distributions of inclination and eccentricity from our functions f(e) and f 1 (i) or f 2 (i). We convert masses to radii using the Seager et al. equation of state and the value for R. The sample of true underlying planetary systems is then observed by randomly choosing a particular plane of observation and calculating the inclinations of all planetary orbits relative to that plane. We also account for the distribution of the planetary lines of nodes with respect to a particular direction chosen for the observer. All planets whose impact parameters are < 1R are then counted as transiting. In order to avoid confusion we will refer to these as tranets in the parlance of Tremaine & Dong (2012). In order to match with the observations we must also reject tranets whose radii are too small to be detectable. In the interests of clarity we will restrict our comparison to the parameter range where more detailed studies (Howard et al. 2012; Youdin 2011) suggest that the completeness is high. These prior studies suggest that the Kepler database is relatively complete for R > 1.5R and P < 200 days and so we adopt this as our conservative observational limit criteria. 3. COMPARISON TO OBSERVATION In order to compare our Monte Carlo model to the observations we adopt the following criteria (following Howard et al. 2012 and Youdin 2011) to define a complete subsample of the publicly released Kepler planet candidate catalog as of Jan 2013 (although we also compare to the sample announced in Batalha et al. (2012) in case there have been any systematic changes in sample selection with time). Since our simulations are for systems around a 1M star we adopt 4100K < T eff < 6100K and logg > 4 in order to approximately match the stellar hosts to the model. We also require Kepler magnitudes < 15 to ensure that the completeness is not reduced because of lower photon statistics and signal-to-noise. We also then require P < 200 days and 1.5R < R < 10R. This leaves us with 463 single tranet systems 79 two tranet systems 26 three tranet systems 7 four tranet systems 1 five tranet system and 1 six tranet

10 Kepler Rocks Fig. 13. The solid points with error bars represent the multiplicity ratios from the Kepler sample as defined in the text. The circles are drawn from the Jan 2013 Kepler catalogue while the triangles are drawn from the previous incarnation (Batalha et al. 2012). Other symbols show samples defined by other authors from Kepler data. The open circles are those from the study of Lissauer et al. (2012) based on the first major sample of multiple tranets. The crosses are for the sample determined by Tremaine & Dong (2012) and the open triangles are those of Fang & Margot (2012) (Figuera et al. 2012 are also consistent although not plotted because they didn t account for five or six tranet systems). The consistency of all such samples indicates that this comparison is relatively robust with regards to sample selection. The relative horizontal offsets are arbitrary and for the purposes of clarity. The solid histogram represents the expected ratios from our Monte Carlo model with a radius enhancement of R = 1.2. The dashed histograms correspond to the same model but with R = 1.0 (long dashes) and R = 1.4 (short dashes). We see that the model robustly reproduces the multiple tranet ratios and is not heavily sensitive to R. In all cases the ratio of singles to doubles is dramatically undercounted. The dotted histogram shows the equivalent of the solid histogram except that we have assumed that another process produces single tranet systems only at a rate equivalent to that which produces the multi tranet systems. Fig. 14. The solid points are the observed distribution of ξ for the sample defined in the text from the Jan 2013 catalog. The open points are defined in the same way but using the earlier sample from Batalha et al. (2012). The dotted histogram represents the distribution from our model with R = 1.2 and using the full eccentricity and inclination distributions. The dashed histogram represents the model result if we keep the same inclination distribution but set all eccentricities to zero. The solid histogram represents an intermediate model in which orbits are circularised only for a < 0.25AU. Note that all points outside the vertical dashed lines are excluded from the fit because of low number counts. system (For the Batalha et al. (2012) catalog the numbers are 463:85:25:2:2:1). 3.1. Mutual Inclinations One of the simplest measures of comparison between data and observation is the ratios of systems of different multiplicity. Figure 13 shows the comparison of observations with the statistics of our Monte Carlo model for different values of the radius enhancement R as well as equivalent measures from several other studies in the literature. We see that the models can reproduce the ratios of triples to doubles and of triples to systems with four or more tranets. We also see that this consistency is robust to uncertainties in R as the predicted ratios don t change much over a range of R consistent with the known planetary mass-radius relation. There is also a consistency of the observed ratios despite different sample selections by different authors (Lissauer et al. 2011b; Tremaine & Dong 2012; Fang & Margot 2012) suggesting that the comparison is also robust with respect to the particular observational cuts. However our models fail to reproduce the observed ratio of multiples to single tranets in the sense that the Monte Carlo model severely underpredicts the number of systems which show a single transiting system. Lissauer et al. (2011b) noted the potential excess of singles and argued that simple geometric considerations suggest that the ratio of singles to doubles should be higher unless there is an unusual correlation or inclination dispersion. To that end we paid special attention to modelling the high inclination portion of the distributions in Figure 11 and to potential correlations in the lines of nodes but neither of these appears to have materially affected the results. Although other studies such as Tremaine & Dong (2012) and Fang & Margot (2012) claim consistency in their modelling of all the data this is because their chosen statistical distributions allow for a substantial population of single transiting systems something not allowed by our more physically derived model. We can of course simply postulate a second unknown process at work that produces only single tranet systems which can then be combined with our model. If we postulate that this second process operates at the same rate as

Hansen & Murray 11 Fig. 15. Solid points represent the period ratio distribution of our observational sample. The solid histogram is the corresponding distribution from the Monte Carlo model with R = 1.2. The dotted histogram is for the same parameters but drawn from the model with an adjusted inner edge ( 3.3). Fig. 16. The solid points are the same as for Figure 15 but the solid histogram now represents only the period ratios in systems where four or more tranets are detected. The dotted histogram represents the period ratios of systems in which two tranets are detected. The distribution of high multiplicity systems better shows the two peaked structure evident in the observations albeit shifted slightly to longer orbital periods. our model we predict a ratio of singles to doubles 0.2 which matches the observations. A second more accurate measure of the mutual inclination distribution is the velocity-normalised transit duration ratio ξ T dur1/p 1/3 1 (17) T dur /P 1/3 2 where T dur is the transit duration and P is the orbital period (Fabrycky et al. 2012). This is an empirical measure of the ratio of the impact parameters for two neighbouring tranets which should scale as the semi-major axis ratio for perfectly aligned systems. Figure 14 shows the distribution of ξ that results from our Monte Carlo model compared to that of the observed sample as defined above. The dotted histogram is for R = 1.2 using the eccentricities and inclinations drawn from the simulated systems. We see that the model distribution captures the wings correctly but underpredicts the peak near ξ 1. This behaviour is the same for all values of R. The dashed histogram shows the model distribution if we use the same R = 1.2 and full inclination distribution but circularise all the orbits. We see that this model captures the observations very well with a χ 2 per degree of freedom of 1.15. Indeed reasonable agreement can be obtained by only circularising orbits within a certain upper semi-major cutoff. The solid histogram shows the ξ distribution for the case in which this limit is 0.20 AU which defines the 2σ level with respect to the fit obtained with full circularisation. We note that the improvement in the fit is solely due to the reduction in the eccentricity we do not assume any change in the mutual inclinations as a result of tidal effects. 3.2. Period Distributions The comparisonofthe model period ratiodistribution to observations shownin Figure15 is not quite asgood as the various measures that depend on inclinations. Some of the structure is reproduced but the simulations underpredict the number of close pairs relative to the number of more widely separated pairs. Furthermore although there is a double peaked structure to the distribution between period ratios of 1 3 the peaks are shifted to larger values than observed. Indeed the entire shape of the distribution appears shifted by 5% to larger values. This suggests that some level of dissipation is required beyond the simple assembly model assumed here. It has been suggested that such enhancements near commensurabilities and the slight shift in the observed values could be the result of tidal evolution (Wu & Lithwick 2012b; Batygin & Morbidelli 2013). We have not implemented this effect although we have included the semi-major axis shift associated with circularisation for values less than 0.20 AU. Some of this mismatch may also be due to poorly quantified selection effects. Steffen (2013) has noted that the period ratio distribution is subtly different between high multiplicity (four or more tranet candidates) and low multiplicity

12 Kepler Rocks (two tranet candidates) systems. In Figure 16 we show the corresponding comparison between the same data as in Figure 15 but now with two distributions drawn from the model in the same spirit as Steffen (2013). We see that the high multiplicity model systems actually do a much better job of reproducing the data than the lower multiplicity systems. The mismatch in Figure 15 is clearly the result of the stronger weighting towards the two tranet systems as in Figure 13 but the structure here suggests that selection effects that favour more high multiplicity systems could potentially improve the comparison as much as moderate radial migration. We note also that the distribution from the two-tranet systems is indeed shifted slightly compared to the higher multiplicity systems as suggested by the comparison by Steffen. We can also compare our simulations to the distribution of absolute periods of detected tranets. This is shown in Figure 17. We see that the shape of the distribution is well reproduced for P >10 days but that the simple model overproduces short period tranets by a factor of three near the peak of the distribution. This is likely a consequence of our rather naive application of a single model realisation to the full population in the sense that our planetary embryo disks all have an inner edge at 0.05 AU and therefore almost always produce a planet with an orbital period < 10 days. If the true underlying model for the planetary initial conditions contains some variation in the location of the inner edge then this peak will be reduced. 3.3. Model Embellishments We have kept our default model simple in order to highlight the robustness of the match to the data but we can also demonstrate that a more comprehensive match to the data is achievable with minor modifications to the model. One mismatch observed in the previous subsection is the overproduction of tranets with orbital periods < 10 days. This is due in large part to the assumption that all protoplanetary disk have the same inner edge. To demonstrate this we consider a modification to our Monte Carlo model in which we adopt an inner edge to the original disk of 0.08 AU in 55% of cases. In the other 45% of cases we adopt the original model. We use the same mass orbital spacing eccentricity and inclination distributions as before. In effect we simply remove those planets that lie inside 0.08 AU in 55% of the cases. We show the results of this model in Figure 17. This modification does not change the quality of the fits to the ξ distribution or the period ratio distribution. Similarly we have not attempted to directly match the observed distribution of radii because of the distinct possibility that there may be significant stochastic variation in the mass of hydrogen accreted/retained by planets. Nevertheless one can match the observed radius distribution with a suitable choice of mass-radius relation. Figure 18 shows the observed distribution of tranetary radii for the sample identified in 3 compared to two distributions derived from our Monte Carlo model. The dotted histogram shows the results for R = 1.2 which captures the median value of the population but does not match the shape. The solid histogram shows the result from the same model except that we now adopt R = 1+0.05(M/M ). This provides a good fit to the tail of the distribution at larger radii and still matches the other observational constraints. This relation bears some resemblance to both the interpolation of L11 and the linear fit of Wu & Lithwick (2012a) although the nature of the fit means it is weighted more by lower mass objects than the higher masses as in the case of Wu & Lithwick. The effective physical model described by this relation is one in which planets have Hydrogen mass fractions < 1% for masses M increasing gradually to roughly 50% at masses 30M. 4. DISCUSSION Our study follows on the heels of several previous analyses of the Kepler multiple tranet candidate systems (Lissauer et al. 2011b; Fabrycky et al. 2012). All of these studies agree that the underlying planetary configuration is best modelled by a population that has an inclination dispersion of only a few degrees and that this is also consistent with the radial velocity sample (Lissauer et al. 2011b; Fabrycky et al. 2012; Figueira et al. 2011; Tremaine & Dong 2012; Fang & Margot 2012). Our analysis affirms these conclusions with the additional benefit of being drawn directly from a physical model rather than a set of ad hoc postulates. Our inability to match the single tranet numbers is also foreshadowed either directly (L11) or by best fit models that require a large fraction of low multiplicity systems (Tremaine & Dong 2012; Fang & Margot 2012). The fact that our underlying model always produces three or more final planets brings this issue into stark contrast. The most convenient explanation is that that the rate of false positives is higher amongst the single tranet sample than the multiples but this would require that current estimates of the false positive rate are severely in error. Lissauer et al. (2012) have argued that the rate of false positives amongst the multiple systems is very low but this necessarily excludes the single candidate systems. Santerne et al. (2012) has claimed that the false positive rate for Jupiter-size candidates is higher ( 35%) than originally estimated (Morton & Johnson 2011) but it is not clear that the same false positive rate will apply to systems with smaller eclipses. Detailed simulations of the Kepler detection process suggests a false positive rate in the range 7%-16% for the size of planets in our sample (Fressin et al. 2013). Thus it seems that we require either an entirely separate contribution of single tranet systems or a process that modifies or edits some fraction of the systems that assemble according to our model. Although we have neglected any effects due to large-scale post-assembly planetary migration in this calculation it could potentially be the alternative contributing process we seek especially if the late-time instability of resonant chains (Terquem & Papaloizou 2007; Ida & Lin 2010) produces primarily systems of low multiplicity. However in that event it is surprising that both processes produce essentially the same orbital period distribution. Figure 19 shows the cumulative distribution for tranets (1.5R < R < 10R ) in single transit systems (open circles) and multiple transit